Sei sulla pagina 1di 9

Electrochimica Acta 48 (2003) 3053 /3061 www.elsevier.

com/locate/electacta

Nanoscale electrodeposition of metals and semiconductors from ionic liquids


W. Freyland *, C.A. Zell, S.Zein El Abedin, F. Endres
Institute of Physical Chemistry, Physical Chemistry of Condensed Matter, University of Karlsruhe, Kaiserstrasse 12, Postfach 6980, D-76128 Karlsruhe, Germany Received 15 October 2002; received in revised form 25 March 2003; accepted 3 April 2003

Abstract The electrocrystallization of Ni, Co and their respective alloys with Al and the electrodeposition of Ge on Au(1 1 1) and Si(1 1 1):H have been studied in the underpotential (UPD) and overpotential (OPD) range. To this end, in situ electrochemical STM and STS measurements have been performed in ionic electrolytes, the room temperature molten salts or ionic liquids AlCl3 / ( [BMIm] ' Cl( and [BMIm] ' PF6 . The larger electrochemical windows of these ionic electrolytes in comparison to aqueous media enables the investigation of electrodeposition of these elements and compounds on a nanometer scale. We present and compare recent results of 2D and 3D phase formation of Ni and Co electrodeposition. Clear differences are observed in the 2D phase formation */Ni monolayer vs. Co island formation */which are discussed in the light of the distinct values of the interfacial free energies of these two metals. In the overpotential (OPD) range, Ni deposition proceeds by a columnar growth of 3D Ni clusters along step edges, whereas Co clusters grow homogeneously at potentials below (/0.17 V vs. Co/Co(II). Electrodeposition of Nix Al1(x and Cox Al1(x is found to be very similar. In both cases, codeposition starts at a potential clearly positive of the Al/ Al(III)-Nernst potential and with increasing Al content smaller grains are observed. The composition of the respective alloy clusters has been probed in situ by STS spectra and it is found that the effective tunneling barriers at different potentials E scale with the cluster composition determined from independent conventional electrochemical and spectroscopic measurements. Finally, we report first investigations of electrodeposition of ultrathin Ge films with varying thickness on Au(1 1 1) and Si(1 1 1):H. Probing the electronic structure of these films by in situ STS spectra a metal /semiconductor transition is indicated with increasing film thickness above 2 /3 nm. This is discussed in comparison with ultrathin Ge films grown by expitaxial vapour deposition. # 2003 Elsevier Science Ltd. All rights reserved.
Keywords: Electrocrystallization; Electrochemical STM and STS; Ni, Co and Ge electrodepositions; Ionic electrolytes

1. Introduction Although electrochemical scanning tunneling microscopy (EC-STM) has been developed only during the last decade, it has already entered modern textbooks in electrochemistry [1]. It is, at present, the only method that can probe the structure at the electrode/electrolyte interface and its change during electrochemical processes in situ with atomic or nanometer resolution. Measuring the differential tunneling conductance (I vs.

* Corresponding author. Tel.: '/49-721-608-2100; fax: '/49-721608-6662. E-mail address: werner.freyland@chemie.uni-karlsruhe.de (W. Freyland).

U spectra) an additional in situ analytical information on the nanometer scale can be obtained by scanning tunneling spectroscopy (EC-STS) [2]. So far, these scanning probe techniques have been applied in a variety of electrochemical studies, but exclusively with aqueous electrolytes, see e.g. [3 /8]. In this way, the electrodeposition of elements and compounds is limited by the decomposition potential of water of /1.2 V. Taking this into account we have started with STM experiments in ionic electrolytes, molten salts at elevated temperatures [9] and room temperature ionic liquids [10,11]. They possess clearly larger electrochemical windows */e.g. up to 7 V for ( [BMIm] 'PF6 at a tungsten electrode [12] */and thus enable electrodeposition of a wider range of metals, alloys and also semiconductors.

0013-4686/03/$ - see front matter # 2003 Elsevier Science Ltd. All rights reserved. doi:10.1016/S0013-4686(03)00378-5

3054

W. Freyland et al. / Electrochimica Acta 48 (2003) 3053 /3061

This aspect is the main objective of the present paper. In the first part, we show representative results of ECSTM measurements with nanometer resolution of Ni and Co electrodeposition on Au(1 1 1) from AlCl3 / [BMIm] 'Cl ( in the underpotential (UPD) and OPD range. The different behaviour of 2D and 3D phase formation of these two metals will be discussed. We then compare the alloy formation of Nix Al1(x and Cox Al1(x on a nanometer scale with emphasis on recent EC-STS measurements of the compositional changes of alloy clusters. Finally, first STM and STS results of the electrodeposition of Ge films on Au(1 1 1) and Si(1 1 1):H are presented focusing on the apparent thickness induced metal /non-metal transition in these ultra thin films.

2. Experimental The room temperature molten salts employed in this ( study, AlCl3 /[BMIm] 'Cl (and [BMIm] 'PF6 , have the following advantages for EC-STM measurements: they have low vapour pressures even at elevated temperatures, have large electrochemical windows, possess a sufficiently high solubility of the respective metal or semiconductor halides and, finally, the AlCl3 containing melts can be varied from Lewis-acidic to -basic characteristics and are particularly suited for Al alloy electrodeposition. Especially from the STM point of view they have shortcomings: they are sensitive to hydrolysis */especially the AlCl3 containing liquids*/ and their specific electrical conductivity is relatively high, of the order of 10 (2 to 10 (1 V(1 cm (1 [13]. This requires a specially developed EC-STM set up similar to the one first described in Ref. [9]. In brief, the electrochemical cell together with the STM scanner and the micrometer screws for coarse positioning are mounted inside a vacuum tight stainless steel housing filled with high purity Ar gas (O2 B/1 ppm, H2O B/2 ppm). The electronics of the scanner is specially sealed so that attack by hydrolysis products like HCl is strongly reduced. Both W and Pt/Ir electrochemically etched STM tips have been used which were electrically insulated by epoxide paint. For the metal electrodeposition experiments reported here the ionic electrolyte AlCl3 /[BMIm] 'Cl ( has been used. Solutions of Ni(II) and Co(II) of (59/0.1) )/10 (3 mol l(1 have been prepared by anodic dissolution of the respective metals at potentiostatic control. Substrates of flame annealed Au(1 1 1) on quartz (Berliner Glas KG) have been employed. Wires of Ni and Co, respectively, were used as counter and reference electrodes, which proved to be very stable. Ge electrodeposition on Au(1 1 1), mica (Molecular Imaging Corp.) and ( Si(1 1 1):H was studied in liquid [BMIm]'PF6 saturated with GeCl4 or GeBr4 at room temperature (see
Fig. 1. Cyclic voltammogram of Ni (upper panel) and Co (lower panel) electrodeposition on Au(1 1 1) from AlCl3 /[BMIm] ' Cl ( (molar ratio 58:42); scan rate: 0.1 V s (1, electrode area0/0.38 cm2.

also [14]). Preparations and fillings of the electrochemical cells and all assemblies of the EC-STM experiments have been performed in an Ar glove box (O2 B/1 ppm, H2O B/2 ppm). The STM and STS measurements have been made with a DI nanoscope E (Veeco Instruments, Mannheim) or an MI controller (Molecular Imaging, Witec GmbH, Ulm). For the electrochemical measurements an MI picostat potentiostat has been used. Further details of sample preparation, synthesis and purification of the ionic liquids, and experimental procedures are given in [15,16].

3. Results and discussion 3.1. Overview of UPD and OPD deposition of Ni and Co on Au(1 1 1) from AlCl3 /[BMIm] 'Cl ( A first overview of the relevant redox processes of the Ni and Co electrodeposition in the potential limits between bulk Au oxidation A and bulk Al deposition F is given by the cyclic voltammograms in Fig. 1. In the UPD range two redox couples are of interest. The couple B/Bl is assigned to Au(1 1 1) step edge oxidation which has been shown in a previous STM study with the same electrolyte [11]. A second redox process C/Cl with an oxidation peak at 0.35 V vs. Ni/Ni(II) is clearly seen in the Ni cyclic voltammogram, but is only weakly indicated in the UPD of Co. This UPD process will be further elucidated and discussed below on the basis of detailed STM images. It is not observed in Ni electrodeposition from an aqueous Watts electrolyte [17]. Bubendorff et al. [18], however, found that UPD of Ni

W. Freyland et al. / Electrochimica Acta 48 (2003) 3053 /3061

3055

analysis of the dissolution peak El in the CVs recorded at different scan rates with different reverse potentials [21] and will be further analyzed with the aid of STM and STS results below. 3.2. 2D phase formation of Ni and Co on Au(1 1 1) In the UPD range of the redox couple C/Cl (Fig. 1) 2D nucleation of Ni and Co on Au(1 1 1) exhibits a clearly distinct behaviour. This is demonstrated by the STM images in Fig. 2. In the case of Ni deposition a complete monolayer is formed within several minutes if the potential is switched from 0.5 to 0.1 V vs. Ni/Ni(II). This is illustrated by the regular hexagonal superstruc ture (Moire pattern) with a lattice constant of 239/1 A ; this is due to and with a modulation amplitude of 0.6 A the lattice misfit between the Au(1 1 1) surface and the Ni monolayer (dNi 0/2.49 A; dAu 0/2.885 A). This interpretation is supported by simultaneously measured current transients which yield an integrated value of the charge of 5309/50 mC cm (2 corresponding to a 0.9 ml of Ni [22]. It is interesting to note that in Ni electrodeposition from an aqueous Watts electrolyte a complete Ni monolayer is only observed in the OPD range at a potential of (/0.1 V vs. Ni/Ni(II) [17]. In contrast to Ni, visible 2D phase formation during Co electrodeposition from the ionic liquid starts only at slightly negative potentials and is characterized by statistically distributed monoatomically high islands (apparent height 1.79/ 0.1 A in comparison to a Au(1 1 1) step edge height of 2.39/0.1 A). Island formation at the first stages of Co electrodeposition has also been reported for aqueous electrolytes [19,23]. In the AlCl3 /[BMIm] 'Cl ( ionic liquid, however, this island growth prevails up to potentials of (/0.15 V vs. Co/Co(II) (Fig. 2b). Details of the kinetics of the island growth are discussed in Ref. [21]. Here, we restrict to a qualitative explanation of the different nucleation mechanisms of Ni and Co as evidenced in Fig. 2. In a simple thermodynamic approach of nucleation a distinction between island and layer growth is determined by the change of the free enthalpy, DG2D, of the formation of a 2D nucleus. For a circular monolayer island of radius R , atomic volume V and lattice constant a, DG2D is given by (see e.g. [24]): G2D 0R2 aV(1 'R2 aV(1=3 '2RE : (1)

Fig. 2. STM images of 2D nucleation of Ni and Co on Au(1 1 1): (a) Moire pattern of a Ni monolayer electrodeposited at E 0/0.11 V vs. Ni/ Ni(II), Etip 0/0.2 V, Itun 0/3 nA; (b) Atomically high Co islands with radii of 2 nm5/R 5/6 nm grown on Au(1 1 1) terraces at a potential E 0/(/0.15 V vs. Co/Co(II), Etip 0/0 V, Itun 0/3 nA; the two larger frizzy islands are Au islands.

on Au(1 1 1) can occur, depending on the nature of anions present in the electrolyte solution. A similar observation has been made for Co electrodeposition on Au(1 1 1) in various electrolytes [19]. In the OPD range, three reduction waves can be distinguished which are assigned as follows: bulk deposition of Ni or Co at D, Al deposition approaching F. The interpretation of wave E by Ni /Al alloy deposition is consistent with previous electrochemical studies [20]. In the case of Co and Co / Al deposition a more complex growth kinetics has to be considered (see [21]). This is indicated by a detailed

Here, the first term describes the energy gain when the film is formed by electrodeposition corresponding to a change in the electrochemical potential Dh , the second term contains the energy cost for the formation of the new interfaces, and sE in the last term is the line tension. In the following, we first consider the role of Ds on the formation of a 2D island/film in equilibrium with the vapour phase (Dh is replaced by Dm0/kT ln(p/p0), with p0 0/saturation pressure). In this case, Ds is given by the

3056

W. Freyland et al. / Electrochimica Acta 48 (2003) 3053 /3061

equilibrium interfacial energies sik according to the Young equation for complete wetting: D 0F;V 'S;F(S;V ; (2)

where F,V 0/film/vapour; S,F 0/substrate/film; and S,V 0/substrate/vapour interface. If Ds 0/0, the substrate is completely wet by the film and layer growth is favoured. On the other hand, if Ds !/0 the substrate is partially wet by the deposit corresponding to island growth. In order to calculate Ds for the interesting case of Ni and Co deposition we have to estimate sAu,Ni and sAu,Co since no experimental or theoretical data are available. In general, the sik scale with the energies of transformation of the two neighbouring phases i and k. Therefore, we may assume that sS,F &/sS,V, sF,V holds. With this assumption and the literature data of the metals of interest here */sCo,V 0/2709 mJ m (2 [25], sNi,V 0/1850 mJ m (2 [26], sAu,V 0/1500 mJ m (2 [27] */we find for Co a clearly positive value for Ds, i.e. island formation is favoured. Taking into account the relatively large uncertainties of the experimental sik data of metals, no clear conclusion can be drawn in the case of Ni deposition. Finally, we have to comment critically, that these considerations are based on the thermodynamic values of the interfacial free energies and do not contain a correction for the finite microscopic thickness of the deposited films. This is crucial for a quantitative understanding of the wetting behaviour of thin films [28]. An important question remains, in how far the above results for the vapour deposition are relevant for the electrodeposition and nucleation of Ni and Co on Au(1 1 1)? In this case, we have to consider that the coordination chemistry of the metal cations in the ionic liquid possibly influences the deposition and growth behaviour. Previous studies on complex formation of transition metals in Lewis-acidic chloroaluminate room temperature molten salts indicate that both Co and Ni form very weakly coordinated [M(AlCl4)3] ( complexes [13,29]. This high chemical similarity of Co and Ni allows to exclude coordination chemistry as an explanation for their different growth behaviour during electrodeposition. For the same reason, we do not think that the first term in Eq. (1) can account for this difference. However, the crystal structures and surface energies of the two metals are clearly different and so the determining contribution in DG2D is that in Ds. For the following reason, we can assume that Ds for the metal/vapour and for the metal/electrolyte interface are comparable in magnitude. Although the absolute values of the sik are strongly reduced going from the vapour to the electrolyte interphase, their difference should vary only slightly. This is indicated by our STS measurements of the effective tunneling barriers of the metal/electrolyte interface (M 0/Ni, Co, Au(1 1 1)), which yield a constant reduction by /4 eV for all three metals relative to

Fig. 3. STM images of 3D cluster formation of Ni and Co on Au(1 1 1): (a) Decoration of Au step edges by columnar Ni clusters at E 0/(/0.03 V, Etip 0/0.02 V, Itun 0/5 nA; (b) Formation of a rough and dense Co cluster lm at E0/(/0.2 V, Etip 0/(/0.25 V, Itun 0/3 nA.

the vacuum level. With the assumption that the interfacial free energies of the metals are reduced by a corresponding constant amount we suggest a Ds !/0 for the Co electrodeposition from the ionic electrolyte, i.e. island formation and growth. This is consistent with our STM observations. 3.3. 3D cluster growth of Ni and Co Passing the Nernst potential a transition from 2D to 3D nucleation occurs for Ni electrodeposition. At slightly negative potentials 3D Ni-clusters start to grow preferentially at the step edges of Au(1 1 1)

W. Freyland et al. / Electrochimica Acta 48 (2003) 3053 /3061

3057

0.2 V vs. Ni/Ni(II) they achieve a height of /30 A and a over a time period of 12 min. When diameter of /100 A the potential is decreased, 3D cluster growth sets in all over the Au(1 1 1) surface [22]. The nucleation kinetics is similar to that reported for Ni cluster growth on an Ag(1 1 1) substrate in aqueous solution [30]. It can be accounted for by a simple model whereby the vertical cluster growth is determined by the strain energy of the structural defects at the substrate, ad metal step edges (see also [15]). 3D growth in Co electrodeposition on Au(1 1 1) from AlCl3 /[BMIm] 'Cl ( starts at potentials below (/0.17 V vs. Co/Co(II). The Co 2D islands grow in size without forming a coherent monolayer before new islands form on top. As a result, a relatively rough film of several monolayers thickness is observed by STM (Fig. 3b). The nucleation kinetics of Co electrodeposition has been studied at various OPDs and follows classical nucleation theory with a critical number of 1 /2 atoms in the potential range (/0.05 V !/h !/(/0.2 V [21]. 3.4. Ni /Al and Co /Al alloy formation Codeposition of metal/aluminum alloys from chloroaluminate melts has been studied by several groups employing conventional electrochemical methods */see e.g. [20,31/33]. In agreement with our results in Fig. 1 they find that Al codeposition sets in at potentials which lie clearly above the Al/Al(III) equilibrium potential, i.e. Ni /Al codeposition starts at /0.3 V vs. Al/Al(III) and that of Co /Al at /0.35 V vs. Al/Al(III) */the AlNernst potential lies at (/0.5 V vs. Ni/Ni(II) and at (/0.6 V vs. Co/Co(II). The compositions and structures of the deposited alloy films have been analyzed by various electrochemical and ex situ structural techniques. For a Nix Al1(x film a detailed analysis of the diffraction peaks suggests a grain size of /10 nm [20]. The in situ STM measurements of this study yield the following trend for the variation of the alloy film morphology of Nix Al1(x and Cox Al1(x . At high x clusters of 1 /2 nm height and 10 nm diameter are observed. Reducing x , the number density of clusters increases fast and they merge and form larger aggregates. Approaching the Al deposition potential, smaller Al-rich clusters of 2/3 nm diameter can be distinguished from the larger alloy aggregates. A typical example is shown by the STM image in Fig. 4 for a Cox Al1(x deposit. An in situ analysis of the composition of the single clusters is possible by measuring the tunneling spectra. For this aim we have measured the I /U curves for different clusters at various deposition potentials and have determined the effective tunneling barrier f according to [2]: I 0const

Fig. 4. Cox Al1(x alloy deposition on Au(1 1 1) from AlCl3 / [BMIm] ' Cl ( melts: (a) STM image of Cox Al1(x alloy clusters deposited at (/0.7 V vs. Co/Co(II); (b) I /U curve of the cluster marked by a cross in Fig. 4a; (c) Variation of the effective tunneling barrier f (E ) with deposition potential E in Cox Al1(x alloy deposition; error bars represent the scattering of f (full symbols) determined from various I /U curves; STS-spectra were taken with W and Pt-Ir tips and at several clusters at otherwise constant conditions of tip distance and potential E .

terraces and at defects of the Ni-monolayer */see the example of Fig. 3a. With increasing OPD these clusters grow in size and number. At a constant potential of (/

U O

exp((A(f(eV)1=2 d) dV ;

(3)

3058

W. Freyland et al. / Electrochimica Acta 48 (2003) 3053 /3061

where A /1.025 eV(1/2 A(1 and d is the tip-cluster distance. The tunneling barrier f is determined by the average workfunctions of tip and substrate. Therefore, if the tip remains the same in a series of measurements with varying potential E , f (E ) measures the variation of the substrate workfunction and thus yields an information of the composition x. A typical I /U curve of Cox Al1(x taken at a cluster marked by a cross in the STM image is given in Fig. 4. The corresponding f(E ) dependence is shown in the lower panel of Fig. 4. The interesting result is that within the deposition range of Cox Al1(x the relative change of f(E ) within experimental errors quantitatively correlates with independent electrochemical measurements of the cluster composition x [20,34]. A similar result we got for Nix Al1(x deposition [15]. Since recording of the tunneling spectra takes only several ms, this is a valuable analytical probe to study compositional changes during electrocrystallization with nanometer resolution.

Fig. 5. Cyclic voltammogram of Ge electrodeposition from ( [BMIm] ' PF6 saturated with GeCl4 on Si(1 1 1):H; electrode area: 0.4 cm2; scan rate: 1 mV s (1.

3.5. Electrodeposition of ultra thin Ge lms and thickness induced metal /non-metal transition Semiconducting nanostructures and ultrathin films have experienced tremendous research interest in recent years because of their potential to develop novel electronic and optoelectronic devices. An example are ultrathin films of Ge whose microscopic and electronic structure has been characterized by a variety of methods in the thickness range from submonolyers to several tens of monolayers */see e.g. [35 /41]. A detailed review of the kinetics and thermodynamics of thin film growth and, in particular, of the effect of strain on the structure of ultrathin Ge films on Si is given by Lagally and coworkers [35]. For films grown at high temperatures, typically at 500 /700 8C and annealed for several hours, Si/Ge interlayer mixing seems to be suppressed by a large kinetic barrier. At low coverage dimer vacancies form and order in a (2n) reconstruction thus reducing the surface strain energy. Pure Ge has a lower surface energy than Si and thus completely wets the substrate up to /3 ml. Above this thickness a transition from 2D to 3D growth (Stranski /Krastanov mode) occurs. It is characterized by the formation of hut-like clusters with canted ends, sometimes of prismatic shape with perfect facet planes [35]. They occur preferentially at lower temperatures (B/500 8C) and are believed to be a metastable intermediate phase. This onset of the transition from 2D to 3D phase formation is clearly identified in STM investigations [35] and is supported by several spectroscopic studies like interference enhanced Raman spectroscopy [38] or UPS [41]. The critical thickness for the transition varies in these different studies which certainly is due to the different temperatures where the films have been grown and studied.

Fig. 6. STM image and I /U tunneling spectrum of a Si(1 1 1):H substrate with 48 miscut recorded at the Si/ionic liquid interface at open circuit potential (1.8 V vs. Ge deposit), Etip 0/2.8 V, Itun 0/1 nA, height scale: 0 /10 A; STS spectra are averaged from measurements at different sites with W and Pt/Ir tips.

In most of these studies, Ge films on various substrates have been prepared by vapour deposition techniques like MBE or CVD at elevated temperatures. In this section, we report new measurements of Ge electrodeposition from an ionic electrolyte on Au(1 1 1) and Si(1 1 1):H at room temperature. In the discussion of the STM and STS results, we focus on the variation of the electronic structure going from very thin to thick deposited Ge films. As an example, a cyclic voltammogram of the ( [BMIm] 'PF6 ionic melt saturated with GeCl4 and measured with a H-terminated Si(1 1 1) electrode is shown in Fig. 5, see also [42]. Measurements have been performed with a Pt quasi reference, the potentials are given vs. Ge bulk deposition. Similar to the interpretation given for the Au(1 1 1) interface [42] we assign the two broad reduction peaks at /0.8 V and (/ 0.3 V vs. Ge/Ge(IV), respectively, to the reduction steps

W. Freyland et al. / Electrochimica Acta 48 (2003) 3053 /3061

3059

Fig. 7. (a) STM image and I /U tunneling spectrum of a /1 nm thick Ge lm electrodeposited from a GeBr4 saturated (/5)/10 (3 mol l (1) ( [BMIm] ' PF6 melt on Au(1 1 1) (see text), E0/0.3 V, Etip 0/0.9 V, Itun 0/1 nA, height scale: 0 /10 A; thickness has been estimated from the z -piezodisplacement after dissolution, same in Fig. 7b and c; all potentials E vs. Ge/Ge(IV). (b) STM image, I /U curve and cluster height prole of a Ge ( lm of 2 /4 nm thickness deposited from a GeCl4 saturated [BMIm]' PF6 melt on Au(1 1 1) (see text), E 0/(/0.05 V, Etip 0/1.05 V, Itun 0/1 nA. (c) STM image and I /U curve of a 4 /6 nm thick Ge lm deposited on a Si(1 1 1):H substrate, E0/(/0.1 V, Etip 0/0.8 V, Itun 0/1 nA.

3060

W. Freyland et al. / Electrochimica Acta 48 (2003) 3053 /3061

Fig. 7 (Continued)

Ge(IV)/Ge(II) and Ge(II) /Ge(0), respectively. Obviously the CV exhibits irreversible behaviour which is not understood in detail. For instance, the role of a finite solubility of Ge in its respective halides is unknown. STM images on Au(1 1 1) and Si(1 1 1):H and in situ I /U curves have been taken at different deposition potentials at room temperature. In Fig. 6, the topography of the Si substrate in contact with the ionic liquid is shown at the open circuit potential of /1.8 V vs. Ge/ Ge(IV). The corresponding I /U spectra measured at different sites of the interface yield a gap energy of 1.19/ 0.1 eV which is in good agreement with the literature value of bulk silicon at room temperature [43]. The variation of the microscopic and electronic structure of Ge deposits of different thickness is illustrated in Fig. 7 by a few STM and STS results representing measurements on Au(1 1 1) and Si(1 1 1):H. In the UPD range, coherent rough Ge films are observed. This is demonstrated in Fig. 7a, for the example, of a 5/1 nm thick ( film deposited from a GeBr4 saturated [BMIm] 'PF6 melt on a Au(1 1 1) substrate. The I /U curve measured on this film exhibits clearly metallic characteristics (Fig. 7a). This metallic behaviour is observed for Ge films on Au(1 1 1) up to a thickness of about 1 nm. Increasing the film thickness a band gap starts to open and the morphology of the Ge films changes significantly. The structure is determined by coherent 3D clusters of trapezoidal shape and with sharp edges which have a height of 2 /4 nm. Fig. 7b gives, as an example, the STM picture and the corresponding I /U spectrum of a Ge deposit on Au(1 1 1). Similar characteristic cluster shapes have been observed for the UPD of Ge on Si(1 1 1):H. In the OPD range, thick Ge films can be

Fig. 8. Variation of the apparent band gap energy with lm thickness d of electrodeposited Ge lms: k, Au substrates; j, Si substrates (see text).

grown and stabilized at a given potential on Au(1 1 1) and Si(1 1 1):H alike. An example is given in Fig. 7c. Taking the band gap of 0.7 eV [43] as a measure, bulk semiconducting properties occur above 10 nm film thickness. The overall change of the electronic structure as a function of the film thickness is presented in Fig. 8. Plotted are the apparent gap energies as determined from the I /U curves of electrodeposited Ge films on Au(1 1 1) and Si(1 1 1):H as a function of the average film thickness d. This plot indicates that a metal/nonmetal transition occurs with increasing film thickness. This is a remarkable observation, since metallic characteristics of Ge so far are only known for the bulk liquid state. The thin films studied here have been grown at a relatively low temperature. At these conditions the

W. Freyland et al. / Electrochimica Acta 48 (2003) 3053 /3061

3061

growth is dominated by kinetics. So it may be that the thin Ge films are far from equilibrium and the metallic like behaviour reflects a metastable glassy state. Further investigations are needed to clarify this hypothesis.

Acknowledgements Financial support of this work by DFG through project FR 299/17 and now through CFN, University of Karlsruhe, and partially by the Fonds der Chemischen Industrie is acknowledged.

References
[1] A.J. Bard, L.R. Faulkner, Electrochemical Methods, Fundamentals and Applications, Wiley, New York, 2001. [2] R.J. Hamers, in: D.A. Bonnell (Ed.), Scanning Tunneling Microscopy and Spectroscopy: Theory, Techniques and Applications, VCH, Weinheim, 1993. [3] O.M. Magnussen, J. Hotlos, R.J. Nichols, D.M. Kolb, R.J. Behm, Phys. Rev. Lett. 64 (1990) 2929. [4] R. Christoph, H. Siegenthaler, H. Rohrer, H. Wiese, Electrochim. Acta 34 (1989) 1011. [5] F.R.F. Fan, A.J. Bard, J. Electrochem. Soc. 136 (1989) 3216. [6] A.A. Gewirth, H. Siegenthaler (Eds.), Nanoscale Probes of the Solid/Liquid Interface, Nato ASI Series E 288, Kluwer Academic Publishers, Dordrecht, 1995. [7] W.J. Lorenz, W. Plieth, Electrochemical Nanotechnology, Wiley /VCH, Weinheim, 1998. [8] E. Budevski, G. Staikov, W.J. Lorenz, Electrochemical Phase Formation and Growth, VCH, Weinheim, 1996. [9] A. Shkurankov, F. Endres, W. Freyland, Rev. Sci. Instrum. 73 (2002) 102. [10] F. Endres, W. Freyland, B. Gilbert, Ber. Bunsen-Ges. Phys. Chem. 101 (1997) 1075. [11] C.A. Zell, F. Endres, W. Freyland, Phys. Chem. Chem. Phys. 1 (4) (1999) 697 /705. [12] P.A.Z. Suarez, V.M. Selbach, J.E.L. Dullius, S. Einloft, C.M.S. Piatnicki, D.S. Azambuja, R.F. de Souza, Electrochim. Acta 42 (1997) 2533. [13] G. Mamantov, A.I. Popov (Eds.), Chemistry of Nonaqueous Solutions, VCH, Weinheim, 1994. [14] F. Endres, S. Zein El Abedin, Phys. Chem. Chem. Phys. 4 (2002) 1640. [15] C.A. Zell, Ph.D. Thesis, Universitat Karlsruhe, 2002.

[16] F. Endres, Habilitation, Universitat Karlsruhe, 2002. [17] F.A. Moller, J. Kintrup, A. Lachenwitzer, O.M. Magnussen, R.J. Behm, Phys. Rev. B 56 (1997) 12506. [18] J.L. Bubendorff, L. Cagnon, V. Costa-Kieling, J.P. Brucker, P. Allongue, Surf. Sci. 384 (1997) L836. [19] L. Cagnon, A. Gundel, T. Devolder, A. Morrone, C. Chappert, J.E. Schmidt, P. Allongue, Appl. Surf. Sci. 164 (2000) 22. [20] W.R. Pitner, C.L. Hussey, G.R. Stafford, J. Electrochem. Soc. 143 (1996) 130. [21] C.A. Zell, W. Freyland, Langmuir, 2003. [22] C.A. Zell, W. Freyland, F. Endres, in: R.W. Berg, H.A. Hjuler (Eds.), Molten Salt Chemistry, vol. 1, Elsevier, Paris, 2000, p. 597. [23] M. Kleinert, H.-F. Waibel, G.E. Engelmann, H. Martin, D.M. Kolb, Electrochim. Acta 46 (2001) 3129. [24] H. Luth, Surfaces and Interfaces of Solid Materials, 3rd ed., Springer, Heidelberg, 1995. [25] L.Z. Mezey, J. Giber, Jpn. J. Appl. Phys. 21 (1982) 1589. [26] J.M. Blakeley, P.S. Maiya, in: J.J. Burke, et al. (Eds.), Surfaces and Interfaces, Syracuse University Press, New York, 1967. [27] J.C. Heraud, J.J. Metois, Acta Metall. 28 (1980) 1789. [28] S. Dietrich, in: C. Domb, J. Lebowitz (Eds.), Phase Transitions and Critical Phenomena, Academic Press, London, 1988. [29] A.J. Dent, K.R. Seddon, T. Welton, J. Chem. Soc., Chem. Commun. (1990) 315. [30] S. Morin, A. Lachenwitzer, O.M. Magnussen, R.J. Behm, Phys. Rev. Lett. 83 (1999) 5066. [31] T.P. Moffat, J. Electrochem. Soc. 141 (1994) 3059. [32] R.T. Carlin, P.C. Trulove, H.C. De Long, J. Electrochem. Soc. 143 (1996) 2747. [33] Q. Zhu, C.L. Hussey, G.R. Stafford, J. Electrochem. Soc. 148 (2001) C88. [34] R.T. Carlin, H.C. De Long, J. Fuller, P.C. Trulove, J. Electrochem. Soc. 145 (1998) 1598. [35] F. Liu, F. Wu, M.G. Lagally, Chem. Rev. 97 (1997) 1045. [36] R.Q. Yu, J. Fortner, S. Chaser, J.S. Lannin, Mater. Res. Soc. Symp. Proc. 206 (1991) 403. [37] S. Ogawa, J. Vac. Sci. Technol. 15 (1978) 363. [38] S. Kanakaraju, A.K. Sood, S. Mohan, J. Appl. Phys. 84 (1998) 5756. [39] P. Castrucci, R. Gunnella, M. Crescenzi, M. Sacchi, G. Dufour, F. Rochet, Phys. Rev. B 60 (1999) 5759. [40] N. Marcovic, C. Christiansen, G. Martinez-Arizala, A.M. Gold mann, Phys Rev. B 65 (2001) 012501. [41] L. Di Gaspare, G. Capellini, E. Cianci, F. Evangelisti, J. Vac. Sci. Technol. 16 (1998) 1721. [42] F. Endres, S. Zein El Abedin, Phys. Chem. Chem. Phys. 4 (2002) 1649. [43] C. Kittel, Introduction to Solid State Physics, 2nd ed., Wiley, New York, 1965.

Potrebbero piacerti anche